keith a mclauchlan-molecular physical chemistry_ a concise introduction-royal society of chemistry...

Upload: csontika

Post on 06-Jul-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    1/140

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    2/140

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    3/140

    Molecular Physical Chemistry

    A Concise Introduction

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    4/140

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    5/140

    Molecular Physical hemistry

    A

    oncise

    Introduction

    K A

    McLauchlan

    University

    o

    Oxford

    U K

    dv ncing

    the chemic l

    sciences

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    6/140

    ISBN 0-85404-619-4

    catalogue record for this book is available from the British Library

    he Royal Society of Chemistry 2004

    ll rights reserved

    Apart fr om air dealing-for the purposes

    o

    research fo r non-commercial purposes or fo r

    private stu dy criticism or review as permitted under the Copyrig ht Designs and Pa tents

    Act

    1988

    and the Copy right and Related Righ ts Regulations 2003 this puhlicution m ay

    not be reproduced stored or transm itted in an y or m or by any means without the prior

    permission in writing of Th e Ro yal Society of Chemistry or in the case of reproduction in

    accordance with th e terms of licences issued by the Cop yrigh t Licensing Age ncy in the U K

    or in accordance with the terms o f t h e licences issued by the appropriate Rep rodu ction

    Rights Organization outside the U K . Enquiries concerning reproduction outside th e terms

    stated here should be sent t o T he Royal Socie ty of Chem istry at the address printed on this

    Page.

    Published by The Royal Society

    of

    Chemistry

    Thomas Graham House Science Park Milton Road

    Cambridge CB4

    OWF

    UK

    Registered Charity Number 207890

    or

    further information see our web site at www.rsc.org

    Typeset by Vision Typesetting Ltd Manchester

    Printed by TJ International Ltd Padstow Cornwall UK

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    7/140

      reface

    To

    write any b ook is an indulgence and my excuse for having do ne

    so

    is

    tha t have tried to provide underg raduates with a text that differs in

    approa ch from any other I kn ow o n its subject ma tter. It is an attempt to

    provide the reader with a n understanding of thermodynamics and to a

    lesser extent) reactions based upo n ato m s and molecules an d their pr op-

    erties rather th an o n one based up on the historical development of these

    subjects. Th is is possible as a result

    of

    innovative experiments performed

    on very small numb ers of ato m s an d molecules that have been performed

    in the last decade o r so

    This book makes no pretence to be a primary source of its subject

    matter. Rather it attempts to give molecular insight into the familiar

    equ ation s of therm odyna mics, for example, and should be read in con-

    junctio n with the excellent Physical Chem istry texts th at a lready exist. It

    is an aid to understanding, no more and n o less. But hope that those

    deterred by the elegant but possibly dry approaches found in other

    books, which develop the subject without the properties of molecules

    considered, will find this m ore to their liking. Th e subjects are imp ortan t

    to the whole understand ing

    o

    Physical Chem istry an d provide the unde r-

    lying philosophical structure th at binds its apparently separate subjects

    together.

    I a m indebted to all my students

    or

    what they have taught me an d for

    the sheer pleasure of kno w ing them . But my g reatest deb t is to Jo an , for

    everything imp ort an t in my life.

    V

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    8/140

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    9/140

    Contents

    Chapter Some Basic Ideas and Exam ples

    1 1 Introduction

    1.2 Energies and Heat Capacities of Atom s

    1.3 Heat Capacities of Diatomic M olecules

    1.4 Spectroscopy an d Qu antisation

    1.5 Summary

    1.6 Fu rthe r Implications from Spectroscopy

    1.7 Nature of Quantised Systems

    1.7.1 Boltzmann Distribu tion

    1.7.2 Two level Systems

    1.7.3 Two level Systems with Degeneracies Gr eate r tha n

    Unity; Halogen Atoms

    1.7.4 Molecular Exam ple: NO gas

    Appendix 1 1 The Eq uipartition Integral

    Appendix 1.2 Term Symbols

    Problems

    Chapter

    2

    Partition Functions

    2.1 Mo lecular Partition Fun ction

    2.2 Boltzman n Distribution

    2.3 Canonical Partit ion Fun ction

    2.4 Sum mary of Partition Fun ctions

    2.5 Evaluation of Molecular Partition Functions

    2.5.1 Electronic Pa rtition Fu nc tion

    2.5.2 Vibrational Partition Fu nction

    2.5.3 Th e Rotational Partit ion Function

    2.5.2.1

    2.5.3.1

    2.5.3.2

    Vibrational Heat Capacity

    of

    a Diatomic Ga s

    Ro tational Heat Capacity of a Diatomic G as

    Ro tational Partition Functions in the Liquid

    State

    1

    2

    5

    10

    14

    14

    16

    18

    19

    22

    24

    26

    26

    28

    3

    30

    33

    36

    39

    39

    40

    41

    44

    46

    51

    53

    vii

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    10/140

    viii

    Contents

    2.5.4

    Translational Partition Function

    54

    2.5.4.1 The Translational Heat Capacity

    55

    Appendix

    2.1

    Units

    57

    Problems 58

    2.6

    Overall Molecular Partition Function for a Diatomic Molecule 56

    Chapter

    3

    Thermodynamics

    3.1 Introduction

    3.2 Entropy

    3.3

    3.4 State Functions

    3.5

    Thermodynamics

    3.5.1 First Law

    Entropy at 0 K and the Third Law

    of

    Thermodynamics

    3.5.1.1

    3.5.2 Second Law

    Thermodynamic Functions and Partition Functions

    Absolute Entropies and the Entropies of

    Molecular Crystals at 0 K

    3.6 Free Energy

    3.7

    3.8 Conclusion

    Problems

    Chapter Applications

    4.1 General Strategy

    4.2 Entropy

    of

    Gases

    4.2.1

    Entropies of Monatomic Gases

    4.2.2 Entropy of Diatomic and Polyatomic Molecules

    Two level Systems; Zeeman Effects and Magnetic Resonance

    4.3.1 Internal Energy of Two level Systems

    4.3.2

    Curie Law

    Pauli Principle and Ortho and Para hydrogen

    4.6.1 Isotope Equilibria

    4.3

    4.4 Intensities of Spectral Lines

    4.5

    4.6 Chemical Equilibria

    4.7 Chemical Reaction

    4.8 Thermal Equilibrium and Temperature

    Problems

    Chapter

    5

    Reactions

    5.1 Introduction

    5.2 Molecular Collisions

    5.2.1 Collision Diameters

    5.2.2 Reactive Collisions

    9

    59

    59

    61

    63

    63

    64

    61

    69

    70

    72

    74

    74

    77

    77

    77

    78

    79

    80

    80

    81

    82

    84

    88

    92

    93

    97

    99

    1 1

    101

    101

    102

    105

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    11/140

    Contents

    ix

    5.3

    Collision Theory

    5.4 Energy Considerations

    5.4.1

    Activation Energy

    5.4.2

    Disposal

    of

    Energy and Energy Distribution in

    Molecules

    5 5 Potential Energy Surfaces

    5 6

    Summary

    Problems

    Answers to Problems

    Some Useful Constants and Relations

    Further Reading

    108

    110

    110

    111

    112

    115

    116

    8

    2

    22

    Subject Index

    23

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    12/140

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    13/140

    CHAPTER

    Some Basic Ideas and Examples

    1.1

    INTRODUCTION

    Phy sical chem istry is widely perceived a s a collection of largely indepen-

    dent topics, few of which appear straightforward. This book aims to

    remove this misconception by basing it securely on the atom s a n d mol-

    ecules that c onstitute matter, an d their properties. W e shall concen trate

    o n just two aspects and we focus mainly o n thermodynam ics, which

    although extremely powerful is one of the least popular subjects with

    students. A briefer account describes how reactions occur. We shall

    nevertheless encounter the major building blocks

    of

    physical chem istry,

    the foun dation s that, if unde rstood , togethe r with their inter-dependence ,

    remove any mystique. These include statistical thermodynamics, ther-

    mody namics and qua ntum theory.

    T he way t ha t physical chem istry is tau gh t tod ay reflects the historical

    process by which und erstan ding w as initially obtain ed. On e subject led to

    another, not necessarily with any underlying philosophical connection

    but largely as a result of what was possible at the time. All experiments

    involved very large numbers of molecules (although when ther-

    modynamics was first formulated the existence of atoms and molecules

    was not generally accepted) and people attempted to decipher what

    hap pen ed a t a mo lecular level from their results. Th is was very indirect.

    Nowadays the existence and properties of atoms and molecules are

    established and experiments can even be performed o n individual atom s

    and molecules. This provides the opportunity for a different way of

    looking at the subject, building from these properties to deduce the

    characteristic behaviour of large collections of them, which is more in

    keeping with how chemistry is taught at school level. Similarly, our

    understanding of how reactions occur has come from observations of

    samples containing -hu ge num bers of m olecules a nd we have tried to

    deduce w hat h app ens a t molecular level from them. Yet it is now possible

    to o bserve reactions between individual p airs of molecules, an d we can

    reverse the procedure and start from these observations to understand

    1

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    14/140

     

    Chapter

    1

    reactions in bulk. It is the object of this boo k to d em ons trate th e possibil-

    ity of a molecular app roac h t o thermody nam ics an d reaction dynam ics.

    It is not intended as a n introduction to these subjects but rathe r is offered

    as an aid to understand ing them, with some prior know ledge assumed.

    W e start with the properties of ato m s and molecules as deduced from

    thermod ynam ic measurem ents an d from spectroscopy. This is, paradoxi-

    cally, the historical approach but it establishes straightaway that the

    properties are directly connected t o the thermod ynam ics an d it is artifi-

    cial to separate the two. B ut once the conne ction is established we show

    how it ca n be exploited to give real insight into various problem s. In this

    chapter we introduce the fact that the energy levels of atoms and mol-

    ecules are quantised an d use som e simple ideas to establish the effective-

    ness of our general approach before proceeding to their origins in the

    second chapter.

    1.2

    ENERGIES A N D HEAT CAPACITIES O F ATOMS

    In the gas phase, ato m s move freely in space a nd frequently collide, at a

    rate tha t de pends up on the pressure of the gas. At atmo spheric pressure

    l o 5

    N

    m-2) and room temperature they move approximately

    100

    mo lecular diam eters between collisions, a t average velocities abo ut eq ual

    to that

    of

    a rifle bullet

    (300

    m

    s-l).

    In elastic collisions some atoms

    effectively stop whilst others gain increased velocity cJ: collisions of

    billiard balls) so that instead of all the a to m s having a single velocity they

    have a wide distribution

    of

    velocities. This is the familiar Maxwell dis-

    tribution (F igure

    1.1)

    ha t results from classical Ne wto nian mechanics. In

    it all velocities are possible but som e are more probable than others. The

    mo st probable velocity depends upon the temperature, as does the width

    of the distribu tion.

    A moving atom

    of

    mass m possesses a kinetic energy of +mu2,where u is

    its velocity. Since in the whole collection of atoms in a gas there is no

    restriction t o th e velocity of a n a to m , there is no restriction to its energy

    either. Using the Maxwell distribu tion (see below), the average energy of

    an ato m can be shown to be

    where

    2

    s the mean square velocity of the atoms in the sample, k is

    Boltzmann’s co nstan t

    k

    =

    R/N,

    where

    R

    is the universal gas constant

    and NA the Avogadro number) and T is the absolute temperature. To

    obta in the to tal energy, E , of a mole of gas we simply multiply by th e to tal

    number

    of

    atoms,

    NA,

    an d obta in

    ( )RT.

    This is the energy due to the

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    15/140

    Some Basic Ideas and Exam ples

    3

    Number of molecules

    1000

    2000

    Velocity/rns-

    Figure 1.1 The Maxwell distribution of velocities of molecules in a gas at 273 and 1273 K .

    A s

    the temperature is increased th e most probable velocity moves t o a higher

    value and the distribution widens, reflecting a greater range o molecular velo-

    cities.

    m otion (translation ) of the a tom s in the gas, the 'translation al energy'.

    Remarkably, although the kinetic energy of an individual atom de-

    pen ds up on its mass, the prediction is tha t the to tal energy of the gas in

    the sample does not.

    It

    seemed so outrageous w hen first made tha t it had

    to be tested, but how? We have calculated the abso lute quantity, E , but

    have no way of measuring it directly. But there is a closely related

    property that we can measure. This is the heat capacity of the system,

    defined as the am ou nt of heat required to raise the tem peratu re of a given

    qu an tity of gas (here

    1

    mole) by 1 K. Different values are ob tain ed if this

    measurement is made keeping the volume of the gas constant (with a heat

    cap acity defined as C,) o r keeping its pressure co nstan t (C,) since in th e

    latter case energy is expended in expanding the gas against external

    pressure. H ere we consider jus t w hat is hap pen ing to the energy

    of

    the gas

    itself, and must use the former. Writing the definition in mathematical

    form, as a partial differential (a differential with respect to just one

    variab le, here T),

    (1.2)

    , =

    g),

    lT),

    ( R

    T,

    = R

    = 12.47

    J

    K - mol-I

    The subscript on the bracket reminds us that we are dealing with a

    constant volume system.

    This is again rem arkable. It says that for all mo natom ic gases, regard-

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    16/140

    4

    Chapter 1

    less of their precise chemical natur e o r mass, the m ola r heat capacity is

    the same, a n d independen t of temp erature. Experiment show s this to be

    correct. For example, He, Ne, Ar and Kr were early shown

    to

    have

    precisely this value over the temperature range 173-873

    K,

    the range

    then investigated.

    It is now wo rth exam ining in more detail where the result tha t the m ean

    energy of a m ona tom ic gas is indepen den t of its natur e com es from. T o

    ob tain this average ove r the whole range of velocities we m ust mu ltiply

    the kinetic energy of a n a to m a t a given velocity by the prob ability tha t it

    has this velocity,

    normalised t o the

    tion (dN /N) is the

    where

    and integrate over the whole velocity distribution

    total number of atoms present. This probability func-

    Maxwell dis tribu tion of velocities.

    Th us the expression for E , clearly and understanda bly, contains the mass,

    m,

    and the velocity, u. Yet due to its mathematical form and since the

    integral is real (an d is evaluated between these up per an d low er limits) its

    value ($kT for m otion in three dimensions, Eq uatio n 1.1) does no t. The

    expression m ay look formidable bu t th e integral has a stand ard form (it is

    a Gaussian function) and its eva luation is straightforward; see Appendix

    1.1. It follows tha t the sam e result is obta ined for any form of energy (no t

    necessarily translational) that can be expressed in the sa me m athem atical

    form,

    +ab2,

    where ‘a’ is a co ns tan t an d

    ‘b’

    is a variable tha t can take any

    value within a Maxwellian distribution. Such a term is known as a

    ‘squared term’.

    F ro m experience, gases are hom ogene ous and possess the sam e prop-

    erties in all three directions in space; for example, the pressure is the sam e

    in all directions. The mo tion of the ga s ato m s in the three pe rpendicular

    Cartesian directions is independent and we say that they have three

    ‘tran slatio na l degrees of freedom ’. Resolving the ve locity in to these direc-

    tions and using P ytha gora s gives, with obvious no tation,

    with an analogous result for their means. In the gas the mean square

    velocities in the three directions are equal. The form of the Maxwell

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    17/140

    Some Basic Ideas and Examples

    5

    distribution we have used is tha t for m otion in three dimensions an d the

    average energy associated with each translational degree of freedom is

    consequently one-third of the value obtained . It then follows that for each

    degree of freedom

    whose energy can be expressed as a squared term

    we

    shou ld expect an average energy of

    i k T

    per atom . This is an im portant

    result of classical physics and is the q uantita tive statem ent of 'the Prin -

    ciple of the Eq uip artitio n of Energy'.

    W e stress that it h as resulted from the Maxwell distribution in which

    there is no restriction of the translational energy that an ato m (o r mol-

    ecule) can possess. F ro m everyday experience this seems em inently rea-

    sonable. We can indeed make a car travel at a continuous range of

    velocities w itho ut restriction (an d luckily perso nal cho ice of how ha rd we

    press the accelerator ra the r th an collisions m ake a w hole range possible if

    we consider a large number of cars ). But is this true of molecules that

    migh t possess othe r sources of energy besides translation ? We sh ould no t

    assume

    so,

    but again put it to experimental test. We shall find later that

    we have to re-examine the case of translation al energy too.

    1.3

    HEAT CAPACITIES

    OF

    DIATOMIC MOLECULES

    Th e heat capacity, C,, of a samp le is directly related to its energy, an d can

    be measured. We expect gaseou s diatom ic molecules, like atom s,

    to

    move

    freely in independent directions in space

    so

    that translational energy

    should confer upo n th e sample a heat capacity

    of R

    = 12.47

    J

    mol-' K- .

    If this was the on ly source of energy tha t m olecules possess then the hea t

    capacity shou ld have this value, and be independe nt of tempe rature. This

    turns ou t to be w rong on both counts. Fo r example, the m easured heat

    capacities (in

    J

    mol-1

    K-') of

    dihydrogen and dichlorine at various

    tempe ratures are given in Ta ble 1.1.

    All these values are substantially greater than expected from transla-

    tional motion. Through the direct relationship between C, and

    E

    this

    implies tha t there must be additional con tributions to the energy of the

    sample. W e no te also tha t for each gas th e value increases with tem pera-

    ture, with a tendency for it to becom e consta nt a t high tem peratures for

    Table

    1.1

    He at capacities ( J mol-'

    K - ' )

    of dihydr ogen and dichlorines at

    diflerent tem peratures

    T ( K )

    Molecule

    298 400 600 800 1000

    20.52

    20.8

    7 21.01 2 1.30

    21.89

    c 2 25.53 26.49

    28.29

    28.89

    29.10

    H2

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    18/140

    6 Chapter

    1

    dichlorine, an d t ha t over the whole range of temp erature s in the table the

    heat capac ity of dichlorine exceeds tha t of dihyd rogen.

    So what forms of energy can a diatom ic molecule have that an ato m

    can not? Th e obv ious physical difference is th at in the m olecule the ce ntre

    of mass is n o longer centred on the atom s. This implies that if there a re

    internal m otion s in the m olecule a s it translates through space these have

    associated energies. The first, and most obvious, possibility is that the

    molecule might rotate. A sample containing rotating molecules might

    therefore possess both translational a n d rotatio nal energy, and we need

    to assess the latter. The simplest, and quite goo d, model for molecular

    rotation is to trea t the diatom ic molecule as a rigid roto r (Figure 1.2)with

    the atom s as point masses

    ( m ,

    and

    m,)

    sepa rated from the centre of mass

    of the molecule by distances r l and r 2 . Classical physics shows the

    rotation al energy to be

    02,

    where

    I

    is its mom ent of inertia an d ci the

    ang ular velocity (measu red in rad

    s-').

    We imm ediately recognise this as a

    'squared term'.

    Rotation might occur about any of three independent axes which in

    general might have different m om ents of inertia, althou gh for a d iatom ic

    molecule two are equal. Taking the bon d as on e axis

    ( z ) ,

    hese are those

    about axes perpendicular to it through the centre of mass and their

    m om ents of inertia a re defined by

    (9

    (ii) (iii)

    Figure

    1.2 Rota tions o j a diatomic molecule. T he atoms are treated as point masses, m , and

    m2,

    with their centres lying along the axis

    of

    the molecule with th e distances

    r l

    and r2 measured hetween these centres and the centre of mass ( C M ) of the

    molecule. The molecule can r otate about three axe s, in the plane of the paper (i),

    about the bond axis

    (ii)

    and out o the plane

    (iii).

    The moments o inertia for

    rotations

    (i)

    and

    (iii)

    are non-zero and equal but the moment of inertia o r rotation

    (ii)

    is zero since there is no perpendicular distance between the point masses and

    the C M along the bond axis. This rotation therefore does not contribute to the

    total rotational energy

    o the molecule.

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    19/140

    Some Basic Ideas and

    Examples

    7

    We note th at the distances are measured in the direction p erpendicular to

    the axes of rota tion , here along th e z-axis. This implies that the mo m en t

    of

    inertia for rotation ab ou t the z-axis is zero because the point masses and

    the centre of mass all lie on a straigh t line, an d n o perpendicular distance

    in the

    x

    or

    y

    directions separates them. We conclude th at only two of the

    three rotational degrees of freedom contribute to the energy of the

    molecule, both throu gh squared terms in the an gular velocity. Using the

    Equipartition Principle we predict that their contribution to the energy

    will be

    2 x i R T J

    mol--'. Th is implies th at , toge ther with the translation al

    con tribution , the tota l energy of the molecule sho uld be

    :RT J

    mol-land

    C,

    should be

    R

    J mol-' K-l. It should no t vary as the tem perature is

    changed.

    This has the value

    20.78 J

    mol-I

    K-',

    which, interestingly and signifi-

    cantly (see later), is very close to the v alue ob served for dihyd rogen a t 350

    K, but Table 1.1 shows

    C,

    to increase with temperature. However, for

    dichlorine it is still much too low at this temperature compared with

    experiment. Once again we conclude that the actual energy is greater

    than we thought, and that the molecule must have another form of

    interna l mo tion associated w ith it. Th is is vibratio n.

    In a vibration the atoms continuously move in and out about their

    average positions (Figure

    1.3).

    As they move outwards the bond is

    stretched, as would be a spring, an d this gen erates a restoring force, which

    i Equilibrium

    position

    + f - -

    Figure 1.3 Vibr ation of a diatomic molecule. T he diagram shows at th e top the (point mass)

    atom s at their distance o closest approach when the y start to move apart again,

    in the centre at their average positions, and at the bo ttom when the bond

    is

    fullest

    stretched and the elasti city

    o

    the bond brings the a tom s back towards each other

    once more. A t the two extrem e positions the a toms are momentarily stationary

    and the molecule possesses the potential energy obtained from stretching

    or

    compressing the bond o nly, but they then start to move, transforming potential

    energy into kinetic energy, a process complete just

    as

    the atoms pass through

    their equilibrium positions.

    I f

    the bond stretching obeys Hooke's Law (the

    restoring for ce generated by moving aw ay fr om the equilibrium position is

    proportional to the distance moved) then Simple Harmonic M otio n results.

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    20/140

    8

    Chapter 1

    if Hooke's Law is obeyed is prop ortion al t o the displacement from the

    equilibrium positions, an d the ato m s return th roug h these positions. In

    this model (also quite good ) the m olecule behaves as a simple harm onic

    oscillator with co ntinually interch ang ing kinetic

    (KE,

    from the m otio n of

    the atoms) and potential (PE, from stretching the bond) energies. The

    total energy is the sum of the two, a nd is conserved in a n isolated gas

    molecule.

    Evib= ( K E + PQ

    At a ny instan t the k inetic energy is given classically by m2 where ,u is the

    'reduced mass' of the molecule (defined as

    m, m2/ (m ,

    +

    m2))

    nd

    v

    is the

    instan taneo us velocity of the a tom s, whilst the potential energy is i k x 2 ,

    where

    k

    is the bo nd force con stant (Hooke's Law co nstan t) an d x is the

    instantaneous displacement from the average position of each atom. A

    diatomic molecule can only vibrate in one way, in the direction of the

    bon d, but because

    of

    having to sum the contributions from bo th forms of

    energy this one degree

    of

    vibrational freedom contributes two squared

    terms to the total energy, throu gh the Equ ipartition Principle, 2 x

    RT J

    m ol? On ce again we have assumed that, in using this Principle, there

    are no limitations on (now) the vibrational energy that a molecule can

    possess.

    Th e total energy of the molecule is, therefore, pred icted t o be the su m of

    the translational (;RT), rotational (RT)and vibrational (RT)contribu-

    tions, giving ZRT

    J

    mol- ' and

    C,

    =

    ZR

    =

    29.1

    J

    mol-'

    K- ,

    greater than

    before but still independent of temperature. This is precisely the value

    obtained experimentally for dichlorine at 1000 K but it is much higher

    tha n tha t of dihydroge n at the same temperature. The heat capacities of

    bo th are still predicted, wrongly, to be indep end ent of temp erature .

    It is now instructive to plot C, against T for a diatomic molecule

    (shown diagrammatically in F igure 1.4).The value jum ps discontinuously

    between the three calculated values, corresponding to translation alone,

    translation plus rotation an d finally translation plus rotation plus vibra-

    tion, over small temp erature ranges (near the characteristic tem peratures

    for rotation an d vibration, Or and

    Ov ib

    Section

    2.5.1).

    These temp eratures

    depend on the precise gas studied, and the changes occur at higher

    tem peratu res for molecules consisting of light ato m s tha n for those t ha t

    contain heavy ones. Only a t the highest tempe ratures are the values those

    predicted by E quipartition. But the contribution from translation alone

    is evident at tem peratu res close to abso lute zero, bu t n ot extremely close

    to it when this contribution falls to zero. In this plot the translational

    contribution is easily recognised through its unique value but which

    of

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    21/140

    Som e Basic Ideas and Examp les

    9

    Figure 1.4

    rotation

    3 5

    2.5

    1.5

    Cvf R

    Vibration

    otation

    Translation

    ii

    293

    0

    ot

    TIK

    Schematic diagram of how

    C,

    for a diatomic molecule varies with temperature. I t

    rises sharply fr om near 0 K and soon reaches 3R/2, expected for translational

    mo tion, where it remains until a higher temperature (near O r , , is reached when

    the rotational degrees offre edo m contribute another

    R

    to th e overall value. T his

    happens well below room temperature (293 K ). fo r all diatomic molecules. A t still

    higher temperatures the vibrational motion eventually contributes another

    R,

    again near a rather well-defined temperature (

    O v i b .

    Th e actual values

    of

    these

    temperatures vary with t he precise molecule concerned, and are lower fo r heavy

    molecules (e.g. Cl ,) than light ones

    (e.g. H 2 ) .

    Th e beginnings of the vibrational

    contribution occur below room temperature for C1, but the u ll contribution is not

    apparent until well above it.

    or vibration con tributes at the lower temp erature is obtainable

    only through further experiment or theory; the rotational contribution

    appea rs at the lower temperature.

    W e conclude tha t molecules exhibit very different behaviour from th at

    we predicted using classical theory an d we m ust exam ine where we might

    have gon e wrong. All the basic equations for rotation al an d v ibrational

    energy are well established in classical physics and ar e no t as sum ptions .

    O n e possibility might be th at the rotationa l an d vibrational energies are

    correctly given classically but that they do not have Maxwellian dis-

    tributions. But also we have made what in the classical world seems a

    wholly unexceptional assu mp tion,

    i.e.

    that there are no restrictions a s to

    the energies a molecule m ight possess in its different degrees of freedom.

    Since the predictions d o not conform t o the experimental observations

    this might be wrong.

    W e speculate that rather th an being able to possess

    any

    values of their rotatio na l an d vib rationa l energies, the molecule may

    be able to possess only specijic values of them . Th is is confirmed directly

    by spectroscopy, see below. W e describe the energies as ‘quantised’. This

    is the basic realisation from which mu ch

    of

    physical chemistry flows.

    Translational energy seems no t to be quantised bu t ac tually is; experi-

    me nts have to be performed a t very close to 0

    K

    to observe this. It is why

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    22/140

    10 Chapter

    1

    C, in Figure

    1.4

    goes to zero a t this tem peratu re. The fact th at different

    diatom ic molecules possess the tw o further forms of energy ab ove cha rac-

    teristic temp erature s th at differ from o ne molecule to the next is an aspect

    of energy-quantised systems that we shall have to understand. But all

    systems behave classically at high enough temperatures, which may,

    however, be below room temp erature. F o r example, all diatomics (save

    H2)exhibit their full rotational contribution below room temperature.

    But only the heaviest molecules exhibit the full vibratio nal co ntributio n

    below very high temp erature s.

    T ha t the lim iting classical beh aviou r is often observed in real systems,

    especially for po lyatom ic molecules, is ultimately w hy we are norm ally

    unaw are of the quantised natu re of the world th at su rroun ds us. But the

    world is quantised in energy and we need t o und erstand an d exploit the

    properties

    of

    matter that this implies. Much of the new technology in

    everyday use depends o n it.

    It

    is fascinating and significant that this conclusion of paramount

    importance was indicated as a possibility through the interpretation of

    classical thermod ynam ic measurements, emphasising tha t a conne ction

    exists between the therm ody nam ics of systems an d the prope rties of the

    individual atom s and molecules tha t com prise them.

    1.4

    SPECTROSCOPY

    AND QUANTISATION

    Q ua nt isa tio n of energy show s itself very directly in the optica l sp ec tra of

    atoms and molecules and initially we consider the electronic spectra of

    atom s. Wh en a sam ple of ato m s is excited in a flame, for example, it emits

    radiation to yield a ‘line spectrum’ (Figu re 1.5). This is in fact a series of

    images of the exit slit in a spectrometer corresponding to a series of

    different discrete frequencies of light emitted by the atoms. If the atom

    behaved classically this would n ot be

    so

    since the electrostatic a ttrac tion

    between the electron an d nucleus w ould accelerate one tow ards the other,

    an d acc ording to classical physical laws the ato m would em it light over a

    continuous frequency range until the electron was annihilated on en-

    cou ntering th e nucleus. Th e explan ation is tha t the energies possible for

    the electrons in a n a tom are themselves quantised, and the frequencies in

    the emission spe ctrum co rrespond to the electron jump ing between levels

    of different energy, acco rding to the B ohr co ndition ,

    where h is Planck’s con stant,

    v

    the frequency of the light and

    E ,

    and

    E ,

    are the energies of tw o of the levels (Fig ure

    1.6).

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    23/140

    Some Basic Ideas and Examples

    11

    Figure 1.5

    Figure

    1.6

    120

    100

    k/nm

    Ultraviolet region of the emission spectrum

    of

    hydrogen atoms (the Lyman

    series). Atoms in the sample have been excited by an electric discharge to a

    number of higher atomic energy levels and they emit light at digerent discrete

    ,frequencies as the electrons return t o the lowest level (principle qua ntum number,

    n

    =

    . Th is displays the quantised nature o the atomic energy levels directly.

    Other series

    of

    lines are observed in the visible and infrared regions

    of

    the

    electromagnetic spectrum, corresponding to electron jumps to diferent lower

    quantised states. The positions

    of

    the lines allow the frequ encie s

    of

    the emitted

    light t o be measured.

    The origin of a line in a spectrum o frequency

    v.

    In a H atom at thermal

    equilibrium with its surroundings the electron is in the lower energy level and it

    can absorb the specific amount of energy hv to ju m p t o a higher energy level. I f

    however, the atom has been excited

    so

    as to put the electron in the upper energy

    level then it can emit the sume energy and return t o the lower level. I n a real atom

    there are many pairs of energy levels between which spectroscopic transitions can

    occur.

    An early trium ph of the Schrod inger eq ua tion was tha t it rationalised

    why the levels are quantised and allowed the energies to be calculated,

    with the difference frequencies agreeing with the observed ones.

    Asso-

    ciated with each level is a mathematical function, the wave function,

    known as an orbital .

    The C, measurements discussed above gave no indication tha t m on-

    atom ic gases could possess electronic energy besides translationa l energy.

    This must m ean that

    ouer

    the

    temperature range studied

    the atoms do not

    hav e any. But we have seen in Fig ure 1.4 th at d ifferent sources of energy

    m ake their contributions a t different tempe ratures and at a high eno ugh

    temp erature there would indeed be an electronic con tribution to the

    C ,

    of atom s. We shall see later th at the crucial factor is the energy sepa ration

    between the qu antised energy levels com pared with th e ‘thermal energy’

    (given by

    k T )

    in the system, and the lower atom ic orbitals are separated

    from each other by large energy gaps. For some atoms, such as the

    halogens (see below) an electronic contribu tion is evident even a t q uite

    low temperatures, bu t this is rathe r unusual.

    In molecules, too, discrete energy levels and molecular orbitals exist

    with differing electronic energies. But associated with each and every

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    24/140

    12 Chapter

    1

    electronic level there is a set of vibrational and rotational ones (Figure

    1.7).

    Th roug h the Bo hr condition spectroscopy allows us to measure the

    energy gaps directly, whilst q ua ntu m mechanics also allows us to calcu-

    late the energy levels, an d the ga ps between them. T his confirms tha t the

    energies molecules possess are quan tised. Molecular sp ectra are norm ally

    observed in a bso rption , with the m olecules absorbing certain frequencies

    from white light flooding through a sample of them. Whereas emission

    spec tra arise w hen electrons fall from higher energy levels into w hich the

    substance has been excited by heat or electricity, absorption spectra

    dep end on species in their lower energy levels jum ping to h igher ones.

    They therefore give inform ation o n th e lowest energy levels of the mo l-

    ecules. Experim entally a huge range of trans ition frequencies is invo lved,

    varying from th e m icrowave (far infrared) to the u ltraviolet regions of the

    electromagnetic spectrum. The former occur between energy levels that

    are closest in energy, those d ue t o rotatio n. At higher frequencies, in the

    infrared, the light has sufficient energy to cause jumps between vibra-

    tional levels, bu t these all have m ore closely spaced associated rota tion al

    ones, an d under certain selection rules changes occur to both the vibra-

    tional an d rotatio nal energies simultaneously. Th e spectra are know n as

    ‘vibration-rotation’ ones (an example is given in F igure

    2.6,

    Chapter 2).

    Finally, at mu ch high er frequencies, in the ultraviolet, electrons can ju m p

    between the vibrational and rotational levels of different electronic en-

    ergy states and the sp ectra reflect sim ultaneous changes in all three types

    ot

    energy.

    Through Equation (1.8) there is a direct relationship between fre-

    quency an d energy difference an d we conclude tha t in terms of the gaps

    between the energy levels

    AE(e1ectronic)>> AE(vibration) > AE(rotation)(and

    >>

    AE(trans1ation)) (1.9)

    This confirms o u r conclusions from

    C,

    measurements (Figure

    1.4).

    At

    0 K

    Cv is zero, but as the temperature is increased translational motion

    rapidly makes the contribution expected from classical physics. At a

    higher temperature the effects of rotation become apparent, and at a

    higher one still, vibration (and at very high temperatures electronic

    energy contributions might appear if dissociation does not take place

    first).

    M oti on s corresponding to th e lower energy gaps make their contribu-

    tions at lower temperatures than those with higher energy gaps.

    It is important to distinguish between the absolute values of the

    various types of energy and the separations between the energy levels.

    Th us the g aps between tran slationa l levels are miniscule but even at very

    low temperature there is a contribution of ( )RT mol-‘ to the energy.

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    25/140

    Some

    Basic

    Ideas and Examples

    13

    Energy

    I

    Electronic

    Vibrational Rotational

    Figure

    1.7

    Schematic diagram o the energy levels o a diatomic molecule showing on ly the

    two lowest electronic levels, which are widely separated in energy. Associated

    with each is a stack o more closely spaced vibrational levels, which in turn

    embrace a series o rotational levels. Under the Simple Harmonic oscillator

    app rox imation, the vibrational levels of each electronic level are equally spaced,

    but the rotational levels predicted by the rigid rotor model are not. T his diagram

    gives an impression

    of

    the relative sizes

    o

    the electronic, vibrational and rota-

    tional energies but o nly a very f e w o f t h e lower energy levels can be shown here

    without loss o clarity . Rea l molecules have fa r more levels. Spectroscopic

    transitions can be excited between the rotational levels alone, between the

    rotational levels in diferent vibrational states and between the sub-levels

    o

    the

    electronic energy sta tes. T he three ty pes occur with very diflerent energies and

    are observed in quite di fer ent regions of the electromagnetic spectrum.

    Th e gaps between electronic energy levels are en orm ous in co mp arison,

    yet electronic energy makes no contribution to the to tal energy of most

    systems a t room temperature.

    A

    com m on er ror is to believe that since the

    energies of molecular electronic levels may be high then the electronic

    energy of the system must be high. If the molecule existed in one of the

    higher levels the energy would indeed be high. But it does not at low

    temperatures, where the molecule is in its lowest electronic level. F o r a

    gas at room tempe rature the translational energy is greatest in magn itude

    followed by the rotation al energy and then the v ibrational one. The o rder

    is exactly reversed from th at

    of

    the energy gaps:

    E(translationa1)

    >

    E(rotationa1)

    >

    E(vibrationa1)

    >>

    E(e1ectronic))

    (1.10)

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    26/140

    14

    Chapter 1

    1.5 SUMMARY

    Atom s an d m olecules may possess several different con tribution s t o their

    total energy b ut each one ca n have only certain discrete quantised values.

    Th e energies associated w ith different m ode s of mo tion differ in m agn i-

    tude, with the g ap s between the energy levels for translation being sm aller

    than those for rotation that, in turn, are smaller than those for vibration.

    The gaps between the electronic energy levels of molecules normally

    vastly exceed all of these. This causes translational motion to occur at

    lower tem peratures than rotation al mo tion, which occurs a t lower tem-

    peratures t ha n vibration (and much lower than electronic excitation). Yet

    we remain u naw are of quantisation in every day life, an d we have som e

    indication already that this is inherently because systems behave classi-

    cally at high en ough tempe ratures. W e need to sha rpen this concept and

    decide w hat a ‘high enough ’ tem pera ture is, an d then we sho uld be able to

    predict the behaviour

    of

    systems from a knowledge of the quantised

    energy levels of the at om s an d m olecules from w hich they are m ade .

    1.6

    FURTHER

    IMPLICATIONS FROM SPECTROSCOPY

    We are so familiar today with spectra that we tend to miss a very

    remarkable fact about them. An atomic spectroscopic transition in ab-

    sorptio n, for example, occurs when an ato m in a specific energy level

    accepts energy from the radiation an d jum ps to ano ther specific level, in

    accordance with the Bohr condition and under various selection rules

    th at limit the transitions th at a re possible. These have been discovered by

    experiment, and can be rationalised using quantum mechanics.

    So

    the

    spec trum of a single at om und ergo ing a single transition is a single line at

    one specific frequency. But w hen we talk abo ut the spec trum of a n ato m

    (a loose term ) we imm ediately think of a w hole family of transition s a t

    different frequencies (Figu re

    1.5).

    Since the electron in on e ato m ca n only

    m ake one jum p between energy levels a t a time it follows that wha t we see

    is th e result of a large number of individual atom s simultaneously absorb -

    ing energy and jump ing t o a whole range of possible q ua ntu m states. Th at

    is, instead of seeing the sp ectrum of a single ato m , we a re observing the

    spectra of a very large number of individual atoms simultaneously.

    We

    say that spectroscopy is an

    ensemble phenomenon,

    meaning precisely tha t

    w hat we observe is the result of wh at

    is

    hap pen ing in the large collection

    of atoms.

    But w hat w ould h ap pen if we were clever eno ugh t o observe a single

    atom over

    a

    long period of time, rath er tha n instantaneously? Following

    the initial abs orptio n of energy from the light beam , the a tom enters a

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    27/140

    Some Basic Ideas and Exam ples

    15

    higher energy level. Let us now postulate that an efficient mechanism

    exists (it does ) for returnin g it t o the lowest level, where it m ight a bs or b a

    different frequency from the incident radiation and attain a second,

    different, higher level. It w ould then retu rn an d the process be repeated

    over a whole cycle of possible transitions covering the total range of

    frequencies. We conclud e th at over infinite time the a to m w ould perform

    all the possible transitions allowed to it and the spectrum of the single

    atom would be identical in appearance to the ensemble one. This has

    been pu t to d irect experimental test in recent years, alth ou gh in mo lecular

    rather than atomic spectroscopy, and found to be correct. It opens the

    possibility of calculating ensem ble behavio ur of a collection of molecules

    from the b ehaviou r over time of a single one, a concept close to the basis

    of using ensembles in statistical therm ody nam ics (see later).

    A noth er unexpected asp ect of spectroscopy lies in the Bo hr cond ition.

    We tend to think of atoms or molecules absorbing energy from an

    incident light beam and jumping to higher energy states, but Equation

    (1.8) does n ot dictate a direction for the energy change to occur. T h a t is,

    whilst an atom in a lower energy state might jump to a higher energy

    state, one in th at s tate might em it energy unde r the influence of the light

    beam , and fall back t o the lower state. In this case the beam would exit

    more intense than it arrived. These processes are known as stimulated

    absorption and stimulated emission respectively. Einstein w as th e first t o

    consider this a nd showed by a simple kinetic argum ent (it is now mo re

    satisfyingly don e using qu an tum mechanics) tha t the abso lute probab ility

    of an upward or a downward transition caused by light

    of

    the correct

    frequency is exactly the sam e. It follows that if there a re m olecules in b oth

    energy levels then the intensity of a spectroscopic line depends on the

    difference between the nu mb er of a tom s th at ab so rb energy from th e light

    beam an d those that emit energy to it, a n d therefore on the

    diflerence

    in

    the populations of the two energy levels. This is further considered in

    Section

    4.4.

    So not only can spectroscopy measure the energy gaps in

    ato m s and m olecules, bu t it indicates this difference in popu lations, to o.

    Since all atoms and molecules at thermal equilibrium with their sur-

    roundings are found experimentally

    to

    exhibit absorption spectra, we

    conclude tha t

    at

    thermal equilibrium the lower energy states are the mo re

    highly populated. Samples in which the a tom s are d eliberately excited to

    higher levels, for example, by an electric discharge through them, have

    their upper levels overpopulated and exhibit emission spectra. This is

    how streets are lighted, using the emission spectrum of sodium. We

    should always remember that systems are not necessarily at thermal

    equilibrium; indeed, equilibrium can be disturbed or avoided in many

    ways

    of

    increasing technical im portance. F o r example, lasers depen d on

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    28/140

    16 Chapter

    1

    deliberately prod ucin g o verp opu lations of higher energy states.

    1.7

    N TURE OF QU NTISED SYSTEMS

    Th e quan tised world is a strange a nd un-instinctive one, complicated by

    the fact that polyatomic molecules possess millions of discrete energy

    levels. But we can understand wh at q uan tisation implies by study ing it a t

    its simplest and we initially consider a system that has just two energy

    levels available to it. Th a t is, one in w hich the energy can not vary w ithout

    restriction, as in the classical physical world, but in which the atoms,

    molecules, nuclei or whatever t ha t com prise the system can individually

    ad op t on e of jus t tw o possible energy states. Such systems actually exist.

    F o r example, the nucleus of the hydrogen atom , the pro ton, is magnetic

    and can interact with an applied magnetic field to affect its energy.

    Experiment shows that it is a q ua ntu m species whose magnetic mo me nt

    can adopt either of two orientations with respect to the field direction,

    rather t ha n the one tha t a classical com pass needle would. In o ne orienta-

    tion the m agnetic m om ent lies alon g the direction of the applied field, an d

    its energy is lowered, whilst in the oth er it o ppo ses it, an d its energy is

    increased. These a re simple experimental facts an d they imply th at appli-

    catio n of the field creates a tw o-level system [F igu re 1.8(i)]. Th is is

    exploited in Nuclear M agnetic Resonance (N M R ) spectroscopy, in which

    transitions are excited between the two levels. A

    different example is

    found in the halogen atoms that behave as though they were two-level

    systems a t low tempe rature.

    Let us be clear what is implied by the existence of the two levels. We

    define the energy of the low er to be

    0,

    an d that of the upper on e

    E

    If we

    have one particle (an atom ,

    a

    nucleus o r w hatever) in its lowest level its

    energy is 0, whereas if, som ehow , we pu t it in to the upp er s tate its energy

    is E It is crucial to o u r understanding of qu antised systems that there is no

    other possibility. Th e particle ca nno t, for example, have a n energy

    of c/2

    or 1.28. This seem s at od ds w ith everyday life in which, up to som e limit,

    systems seem to be ab le to possess any energy. F o r example, we can h eat a

    kettle to any temperature below the boiling point of water. We need

    somehow to reconcile this difference with the classical world since we

    know that a t atomic o r molecular level

    all

    energies are quan tised.

    Th e secret once mo re lies in the fact th at in the exp eriments we usually

    perform we do not study individual particles but rather collections of

    them in which they may be dis tributed between the two energy states. W e

    sta rt by simply add ing a seco nd [Figure 1.8(ii)]. N ow bo th m ay be in the

    sam e energy level, to give a tot al energy of 0 o r 2c or on e may be in one

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    29/140

    Some Basic Ideas and Examples

    17

    Magnetic Field

    - I

    0

    (ii)

    Figure 1.8

    (i)

    A two-level system results when hydrogen nuclei are placed inside a magnetic

    field. Some protons align with their magnetic moments along the applied j e l d ,

    and some against it, resulting in two difSerent energy states. At thermal equilib-

    rium there are more nuclei in the lower energy state than in the upper one.

    Nuclear Magnetic Resonance Spectroscopy consists in causing a transition

    between the two.

    (ii)

    If just two particles enter

    a

    two-level system their total

    energies can be

    0 , 2 ~r E,

    but i ft h e y enter the levels with equal probability then

    the latter can be obtained in two way s, unlike 0 and 2~ which can be obtained in

    jus t one.

    level and the other in the other, giving a total of

    E .

    Now consider the

    average energy of the two. F o r this latter case this is &/2,

    so

    tha t we now

    have an energy tha t is not one of the quan tised values. Th e total energy is

    simply the n um ber

    of

    particles times the average value

    (2

    x

    ~ / 2 ) , nd in

    this case has a value equal to th at

    of

    one

    of

    the qua ntised levels. But this is

    rarely true as we increase the num ber

    of

    particles in the system. Co nside r

    one in which there are

    ( m +

    n) particles divided between th e tw o energy

    states, with

    m

    in the lower energy one. N ow the total energy is

    nE,

    which

    can take a range

    of

    values determ ined by n, an d the average energy

    (1.11)

    which is determined by the values of m and n. W hereas the energies of the

    individual particles have discrete, quantised, magnitudes, the average

    energy, and the total energy, have no such restrictions and can take a

    wide range

    of

    values. This lies at the heart of why systems containing

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    30/140

    18 Chapter 1

    species with q uantised energy levels nevertheless display classical behav-

    iour und er m ost c onditions. These energies clearly depend on how the

    particles are distributed over the two energy states, tha t is on the num bers

    m

    and

    n.

    This simple example leads to a further important insight. If we can

    distinguish between the particles,

    i.e.

    know which is which, then the

    situations with

    0

    o r 2~ in total energy can each be obtained in just one

    way. But an energy of E is obta ined if either is in the low er level an d th e

    oth er in the upp er o ne, in tw o w ays. If b oth levels are equ ally likely to be

    populated this energy is twice as likely to occur as either of the others.

    Generalising, this implies that som e population distributions a nd some

    energy values ar e m ore likely to occur tha n others. This is

    a

    conclusion of

    mo men tous im portance. However, we stress tha t it depend s on the likeli-

    hood

    of

    pop ulating each level being inherently equal.

    We shall now consider systems containing

    a

    large number of, e.g.

    molecules, which may exist in any of a large number of energy levels,

    before returnin g t o the two-level system.

    1.7.1

    Boltzmann Distribution

    Within chemistry we habitually deal with systems that contain a very

    large numb er ( N )of molecules an d we consider how these are distributed

    between their numerous energy levels when the systems are in thermal

    equilibrium with their surroundings at temperature T. For example,

    1

    mole of gas at 1 bar pressure contains N ,

    (6.022

    x molecules. W e

    need the q ua nt um analo gue of the M axwell distribution of energies. It is

    given by the Boltzmann distribution, which we shall state and use here

    before deriving it in Chapter

    2.

    It is a statistical law that applies to a

    constant number

    of

    independent non-interacting molecules in

    a

    fixed

    volume, an d is subject to the to tal energy

    of

    the system being consta nt

    (the system is isolated), an d the several ways

    of

    obta ining this energy by

    distributing the molecules between the quantised levels being equally

    likely. That is, it does not matter which particular molecules are in

    specific energy states provided tha t the total energy is con stant. The

    distributio n is

    (1.12)

    where ni s the number of molecules in a level of energy ii and

    g i

    s the

    degeneracy

    of

    that level (the number

    of

    states

    of

    equal energy, for

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    31/140

    Some Basic I d e m and Example s

    19

    example, the three p o rbitals

    of

    a

    H

    ato m have the same energy and

    so for

    this

    g i

    =

    3).

    The denominator includes summation over all the energy

    levels of the molecules. This equation results from statistical theory

    applied t o a very large num ber of molecules, und er which condition s on e

    particular distribution of the molecules between the quantised levels

    becomes

    so

    mu ch m ore likely than the rest that it alone need be consider-

    ed. This is the ultimate extension of ou r conclusion concerning just two

    levels.

    O n first encounter, the B oltzman n d istribution looks formidable, es-

    pecially because it apparently involves summation over the millions of

    energy levels present in po lyatom ic molecules. But we now return to the

    two-level system to discover the circumstances where this is only an

    ap pa ren t difficulty.

    1.7.2 Two-level Systems

    Consider a two-level system in which there are no particles (atoms,

    molecules or wh atever) in the lower level an d n l in the upper on e so that

    the total number N

    =

    (no+ nl). At therm al equilibrium the Boltzmann

    distribution tells us that these are related th roug h

    Rea rrangem ent yields

    (1.13)

    (1.14)

    In the simplest case the degeneracy of each state is 1, e.g. for protons

    inside a magnetic field. H ere the ratio of the po pulatio ns dep end s directly

    an d solely on the value

    of

    the dimensionless expon ent (@T) ha t varies as

    the temp erature is change d, being a fixed characteristic of the system.

    The denom inator,

    kT,

    is kno wn as the 'thermal energy'

    of

    the system. This

    energy is always freely available to us in systems a t therm al equ ilibrium

    with their su rroundings a nd, indeed, we canno t avoid it without decreas-

    ing the temp erature to

    0

    K.

    Th is gives us a simple physical picture. The

    therm al energy is wh at a system possesses by virtue

    of

    the motion

    of

    the

    particles th at com prise it, an d we see th at it is closely related, for exam ple,

    to the mean thermal energy due to translation +kT, above. But the

    distribution tells us th at in q uantised systems we must c om pare

    kT

    to

    rather than

    2

    imes it. If

    kT

    is mu ch less than

    E

    we do n ot have the energy

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    32/140

    20 Chapter 1

    to raise the particle from its low er energy level to the higher on e, but as

    the temp erature is increased it becomes possible t o d o

    so.

    So much for the basic picture; now let us investigate the distribution

    semi-quantitatively. At low temperatures

    kT

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    33/140

    Some

    Busic

    Ideas

    and

    Examples

    21

    T T T

    0)

    (ii) (iii)

    Figure 1 9 (i) I n a two-level sys tem all the population is in the lower level at absolute zero

    but t he number in the upper level

    (n , )

    grows as the tem perature is increased. I t

    does not grow indefinitely, however, but reaches an asymptotic value at high

    temperature.

    I f

    the levels are singly-degenerate, half of the total number

    o

    molecules is then in each level.

    (ii)

    Th e total energy

    of

    the sys tem is the product

    o

    the population o the upper level times its energy (see te xt ) and

    so

    it varies with

    temperature exactly as does n l . Th e energy also reaches an asym ptotic value at

    high temperature. (iii) Variation o C,with temperature is given by the difleren-

    tial with respect t o temperature

    of

    the energy curve,

    (ii)

    It starts

    from

    zero , goes

    through a ma xim um at the point o inflexion of the energy curve, and returns to

    zero a t high temperature.

    This is a simple multiple

    of

    y1

    so

    that the variation of energy with

    temperature has exactly the same form as the variation of n , [Figure

    1.9(ii)]. This is astonishing. It show s th at as th e tem peratu re is increased

    E

    does not increase continually but again tends to a n asym ptote. T o check

    this we must again turn to a calculation a nd m easurement

    of

    C,, which is

    obtained over the temperature range simply by differentiating Figure

    1.9(ii). [ C v = (aE/aT),]. Th is is shown in F igu re 1.9(iii). It predicts, what

    wou ld be very stran ge in classical physics, th at the heat cap acity increases

    through a maximum and then falls

    to

    zero as the temperature is in-

    creased, a n d is wh at is observed expe rimentally.

    This precise behaviour is unique to the two-level system. But the

    argum ents we have used are no t. Th us the B oltzmann distribution in any

    system, in which an y particle m ay exist in an y on e

    of

    a number of discrete

    energy levels, always contains exponential terms in which quantised

    energies are c om pared with kT, an d it is the values of these exp onentials

    th at largely determ ine level pop ulation s a nd all the physical prope rties of

    the sample. That is, they all depend upon the ratio ( /kT).his simple

    realisation gives prob ably the most im po rtan t insight in to physical chem -

    istry.

    An exam ple is seen if we extend th e arg um ent used abo ve to calculate

    the energy of the two-level system t o on e with m any levels. W e realise tha t

    the particles are distribu ted am ong st the levels each

    of

    which has its own

    characteristic qu antised energy, ii nd we must sum the energies

    of

    them

    all. It follows tha t

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    34/140

    22

    E = E n i s ,

    1

    Chapter

    I

    (1.19)

    This seemingly obvious statem ent h as astounding implications when we

    realise what we have done. It says that if we know the values of the

    qua ntised energies, which we can determ ine from spectroscopy o r calcu-

    late using qua ntu m m echanics, then we can calculate a thermody nam ic

    property. It establishes a direct relationship between the properties of

    individual atoms and molecules and the thermodynamic properties of

    samples made u p of large numb ers of them, an d it is one of the funda m en-

    tal equa tions of statistical thermody nam ics. W e shall return t o it later.

    1.7.3

    Two-level Systems

    with

    Degeneracies Greater than Unity;

    Halogen Atoms

    At roo m tempe rature the electrons in ato m s are found exclusively in their

    lowest orbitals. Th is is because the highe r orbitals a re greatly separa ted in

    energy from them

    so

    that (&/kT)

    >

    1.Th is is true of the halogens, bu t w ith

    these the lowest level is split into tw o by spin-orbit coupling, which

    is

    smallest in fluorine and largest in iodine (Figure

    1.10).

    Such coupling

    results because motion of the electrically charged electron around the

    nucleus in a p -orbital causes a m agnetic field there th at is experienced by

    the electron itself. Ho wev er, the electron possesses spin angu lar m om en-

    tum tha t causes it to have a q uite separate magnetic mo me nt. As with the

    proton in an external magnetic field the quantised electron magnetic

    moment can adopt just two orientations inside the field due to orbital

    mo tion, an d two energy levels result. Experiment shows tha t the energy

    sepa ration between them is very low com pared with the energies sepa rat-

    ing the orbitals

    so

    that the halogens behave as if they were two-level

    systems a t n orma l temperatures.

    2p3 2

    0

    g = 4

    Figure 1.10 The lowest electronic energy level of the halogen atoms is split into two by

    spin-orbit coupling, the interaction between th e magnetic moments due jirs tly

    to the orbital motion

    of

    the electron in its p-orbital and secondly due to its

    intrinsic spin. Inpuorine the splitting in energy is of the order

    of

    k T a t room

    temperature and both levels are populated. The next lowest electronic level is

    comparaticely very high in energy (energy

    >>

    k T ) a n d is completely un-

    populated at room tem perature. A t thi s temperature the atoms behaue as though

    they are two-level system s, but the t wo levels have difer en t degeneracies.

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    35/140

    Some Basic

    Ideas and Examples

    23

    We take F as an example. Its electron configuration is ls22s22p5 o

    tha t it has a single unpaired electron in a p-orbital. The states that result

    from spin-orbit coupling can be calculated simply using Term Sym bols

    (Appendix 1.2) tha t show the gro und state to split into

    2P1,2

    nd

    2 P 3 , 2

    components. The subscript indicates the total angular m omentum qu an-

    tum num ber of the state, J , and according to H und's rule the state with

    the highest J , 4, ies lowest in energy w hen a n electron shell is over half

    full. How ever, the Term Sym bol has an oth er crucial piece of inform ation

    encoded in it since the degeneracy of a sta te is (2 J

    + l ) ,

    an d we have seen

    that the degeneracy enters the Boltzmann distribution. For the upper

    state g,

    =

    2 whilst for the lower o ne

    go =

    4 (Fig ure 1.10)so that

    (1.20)

    Th e expo nential term changes with tempe rature exactly as before, tend-

    ing to unity as

    T

    -

    00.

    In conseq uence the asym ptotic value of the ratio is

    no longer 1 but 0.5. Th at is, at high tempe ratures one-third of the ato m s

    are in the upper state compared with the half obtained when the levels

    have equal degeneracies. Had the state with

    J

    =

    been the upper one

    then tw o-thirds of the atom s would have been in the upper state at the

    higher temperatures. Simply put, a t high temp eratures, a state of degener-

    acy g can hold

    g

    times m ore atom s than one of degeneracy one -t h e states

    behave a s thoug h they were buckets. Deg eneracy has a significant effect

    on level populations.

    Th rou gh the direct relationship between energy an d heat c apacity it is

    clear tha t the halogen ato m s have C, values tha t reflect their ab ility to

    accept electronic energy within these split ground state levels besides

    possessing trans lationa l energy.

    It rem ains to pu t in som e values to see how significant this is. No tably,

    the exponent always appears as a ratio so, provided that we express

    num erator a nd deno mina tor in the sam e units, i t does not ma tter what

    the un its are. Spectroscopists measure

    E

    using experimen tally convenient

    reciprocal wavelength units, den oted an d usually quo ted in cm-' (1

    cm

    =

    m). These are directly related to energy th roug h the relations

    E

    = hv

    and c = v / l o r c =

    vV,

    where

    v ,

    1 and c are respectively the

    frequency, wave length and velocity of the light. It follows th at

    E

    =

    h c b .

    But rather than calculating this each time it is convenient to calculate

    kT/hc in cm-I an d to use the measurem ent un its. (A discussion o n units is

    provided in Appendix 2.1, Ch apter 2). F o r F,

    =

    401 cm-l, whilst kT/hc

    a t 298 K (room temperature)

    =

    207.2

    ern- ,

    so

    that

    E/kT

    =

    1.935, and

    e - ~ / k T= 0.144. T o work ou t the electronic contributio n to the energy

    of

    1

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    36/140

    24 Chapter

    1

    mole of F atoms at this temperature we first return to the Boltzmann

    distribution to ob tain the num ber of atom s in the upper state:

    g

    l e

    l k T

    2

    x

    0.144

    = 0.067

    1

    N A

    g o e WkT + y lecE ikT 4

    +

    2

    x

    0.144

    (1.21)

    where, as usual, we have defined the lower state t o have zero energy. Th e

    total electronic energy per mole at this temperature is

    E = n , & N , =

    0.067

    x

    401N, cm-' mol-', wh ich is 328

    J

    mol-l. Th us the

    presence of the low-lying sta te increases the tot al energy of the

    system from the pure translation value of 12.47 x 298

    =

    3716 J rno1-I

    (recall

    C

    =

    12.47

    J

    mo1-I K-' for tran slation) by roughly 9 % . This

    increases rapidly with tem pera ture.

    The electronic heat capacity is given by

    C,

    = (dE/dT) = &(dn,/dT),

    where the latter is obtained by differentiating the previous eq ua tion .

    1.7.4

    A

    Molecular Example:N O

    Gas

    NO

    is the only simple diatomic m olecule th at contains a single unpaired

    electron and it is in a

    n*

    orbital, implying that it possesses one unit of

    orbital angular momentum about the bond axis, yielding a magnetic

    mo m ent. Magn etic interaction with the electron spin magnetic mo me nt

    once mo re results in spin-orbit coupling (Figure 1.11) and the gro und

    state is split into two, giving

    2111.,2

    and 2113,2 tates, with now the former

    the lower in energy. In diatomic molecules, as opposed to atoms, the

    degeneracies cannot be assessed from thesesymbols, but each is doubly

    degenerate

    (go

    =

    g1 =

    2). Th e higher sta te lies 121 cm-' abo ve the lower

    so

    that a t room temperature

    e /kT

    .58 and

    (1.22)

    showing that over one-third of the molecules are in it at room temp era-

    ture. As before, we could wo rk ou t wh at this implies as a mo lar contribu-

    tion to the tota l energy of the system but it is obviously appreciable, an d

    so

    is the effect on the heat capacity. But, as with all light diatomic

    molecules, there is n o co ntribu tion to the heat capacity a t this tempera-

    ture from vibratio nal m otion, since the first excited vibration al level is too

    far removed in energy from the g roun d state to be occupied.

    Clearly, it is simple t o calculate the e lectronic energy an d hea t capacity

    of F and

    NO.

    But our calculations need not be restricted to these

    properties. A sample

    of NO

    a t room tempe rature is found to be magnetic

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    37/140

    + - -

    t P

    b Ps

    -

    - -

    (ii)

    Figure

    1 11

    I n

    N O

    there is an unpaired electron in a n-orbital w ith orbital angular momen-

    tum (with quantum number = 1) about the bond a xis; the vector representing

    this is therefore drawn along this axis. (i) Relative to this the spin angular

    mome ntum vector o the electron s =4 can lie either parallel to it, to give an

    overall momentum of   ,or antiparallel to it 4, eading to I IS l 2nd 211112 tates.

    (ii)

    These motions

    o

    the negatively charged electron produce magnetic mo-

    men ts, also along th e axi s but in the opposite direction to the angular momen-

    tum vectors, and experiment shows that the magnetic moment due to spin

    motion is (almost) equal to that due t o orbital motion. In the for me r state th e

    moments add to make the state magnetic but in the latter the moments are

    opposed and the state is not magnetic. I n the molecule, as opposed t o the atom ,

    each state is doubly degenerate g =

    2)

    and, with the next lowest molecular

    orbital well removed in energy from the lowest, N O behaves as a two-level

    system e xac tly as shown in Figure 1.9.

    (actually param agnetic). At first sight this seems unsurprising in a mo l-

    ecule tha t co ntains an unpaired electron, since electrons are m agnetic, but

    we m ust rememb er th at in spin-orbit cou pling we have discovered the

    influence of a second m agn etic field within th e m olecule, due to orb ital

    m otion. W e have therefore to consider the resultant m agnetic field of the

    two ra ther tha n just tha t of electron spin. By qua ntu m laws these can lie

    only parallel (the 21-1312 tate) or antiparallel (the 2111j2tate) to each

    oth er alon g the molecular axis. In the former case the m agnetic mo me nts

    re-enforce each other, and in the latter they are opposed. It was dis-

    covered experimentally that the m agnetic mom ent d ue t o the spin m otion

    is

    almo st exactly equal (to 0.11 ) to tha t due to orbital motion so that the

    two cancel in the

    2111,2

    tate, m akin g it essentially non-m agnetic. Since

    this is the lower energy sta te all the molecules w ould be in it a t sufficiently

    low temperature and the sample would not be appreciably magnetic.

    Th at a roo m tem perature sam ple is magnetic results from therma l popu-

    lation of the up per state.

    25

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    38/140

    26

    Chapter 1

    A

    measure of the magnetism of a bulk sample is its susceptibility

    x,

    Greek chi) which is again straightforward to calculate. Calling the mag-

    netic m om ent of

    a

    molecule in the 2113 2tate

    p,

    it is given by

    and n,,z per mo le is obtained from the B oltzmann distribution as before

    (but remember

    go = g1

    here). Th e tem peratu re dep endence of the suscep-

    tibility ha s exactly the sa m e m athem atical form as d oes the pop ula tion of

    the upper state, an d th e energy of the system. It therefore increases from

    (near) zero at

    0 K

    an d rises to an asym ptotic value. We see how powerful

    some rathe r s traightforward ideas are in calculating the physical pro per-

    ties of collections of ato m s and molecules.

    APPENDIX 1.1

    THE

    EQUIPARTITION INTEGRAL

    From Equations

    (1.3)

    and

    (1.4),

    Since the stan dard integral

    (1.24)

    (1.25)

    we find th at th e terms in m cancel and

    mu2

    = SkT

    (1.26)

    This result obviously holds for an y ‘squared term’ whose energy distribu -

    tion

    is

    given by the Maxwell equa tion.

    APPENDIX 1.2 TERM SYMBO LS

    A Term Symbol provides a straightforward means for assessing what

    states of an ato m exist as a result of spin-orbit coupling from kn owledg e

    of the electronic structure

    of

    the atom . In general an atom possesses many

    electrons and coupling occurs between their spin and orbital motions

    according to definite rules. For light atoms (those with low atomic

    numbers) Russell-Saunders coupling decrees that the spin and orbital

    angular m om enta sum according t o the rules:

    s csi

    I

    (1.27)

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    39/140

    Some

    Basic Ideas

    and

    Examples

    27

    L = p i

    I

    (1.28)

    where

    S

    is the vector sum of all the individual spin vectors

    of

    the

    electrons,

    si

    nd likewise for

    L ;

    the qua ntum unit of angular mom entum

    (h/2n) is omitted by con vention. Th e total angular mo me ntum resulting

    from coupling between the magnetic moments due to these resultant

    m om enta is then given by

    J = L + S

    (1.29)

    A further convention is to write all these quantities in terms of the

    qua ntum numbers (scalars) rather than actual values

    of

    the angular

    mom enta, despite having to remember tha t vector addition is involved.

    Th e result is expressed in the T erm Sym bol

    2 s + lLJ (1.30)

    Orbital angular momentum quantum numbers of individual electrons

    are given letter symbo ls. Th us

    s,

    p, d and refer to values of

    0,

    1 , 2 an d

    3

    respectively. W hen the vector ad dition has tak en place capital

    S,

    P,

    D

    and

    F symbols are used for the correspo nding tota l

    L

    values.

    In filled electron shells individual si values cancel, as do the corre-

    sponding

    i

    ones, an d we need consider only the u npa ired electrons. In the

    F atom , with one unp aired electron in a 2p-orbital,

    S = s i =

    and L=

    i =

    1, which is a

    P

    state. By the general laws of qua ntu m mechanics the

    two vec tors can lie only parallel o r antipara llel to each other, yielding jus t

    two possible values of

    J ,

    4and

    .

    Their Term Sym bols are 2P3,2nd 2Pli2

    respectively, since (2 s + 1)= 2. They are described as ‘doublet P three

    halves’ an d ‘doublet P half’ states. Their energy separation depends on

    the spin-orbit cou pling con stant, which differs between different halogen

    atoms . Fr om general qua ntu m m echanical principles, the degeneracy of

    each state is given by ( 2 J + 1).

    W ith diatom ic molecules a sim ilar con ven tion is used, except that the

    total angular momenta are summarised in Greek alphabet symbols,

    C

    (sigma), Il (pi), an d

    A

    (delta) corresponding to S, P and D in atoms. The

    basic symbo l becomes

    2c

    ‘A where s analogous to

    S

    and A (lambda)

    to

    L,

    an d spin-orbit coupling ma y again occur between the two m om en-

    ta . Fo r NO with its single unp aired e lectron in a

    n*

    orbital (A = 1, a ll

    state) the resultant states are, consequently, and

    2113/2.

    ut these

    both have degeneracies

    of

    2, no t wh at we would expect for atom s. They

    ar e referred t o a s ‘doublet Pi half’ an d ‘doublet Pi three halves’ states.

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    40/140

    28 Chapter 1

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    41/140

    Some Basic Ideas and Examples

    29

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    42/140

    CHAPTER 2

    Partition Functions

    Chapter 1established that atom s a nd molecules possess quantised energy

    levels and that the energy gaps between them can be measured directly

    using optical spectroscopy . If the energy of the lowest state is know n th en

    this allows absolute values for the energies of the systems to be ascer-

    tained. In all cases but on e, the energy du e to v ibration in a m olecule (see

    below), the lowest energy is defined to be ze ro. In addition , assuming the

    Boltzmann distribution for the populations of the energy levels in systems

    a t thermal equilibrium with their surrounding s allowed us to calculate

    the physical properties of some two-level systems. We could take the

    agreem ent between results

    so

    calculated a n d experiment as evidence that

    the distribu tion is correc t. But th e theoretical d erivation

    of

    the d istribu-

    tion gives insight into the conditions under which it is valid, and we

    return to this below. Meanwhile we continue to demonstrate that the

    simple ap pro ach we have used w ith two-level systems can be generalised

    to ones con taining an y num bers of levels.

    2.1 MOLECULAR PARTITION FUNCTION

    Th e energy

    of

    any quantised system can be ob tained, as stated above, by

    summ ing over the popu lations of the individual s tates multiplied

    by

    the

    energies of those states:

    This is, though, an impractical formula for estimating E for systems,

    including molecules, in which the number

    of

    energy levels may be very

    large indeed. First, it involves a su m m ation over term s in each of these

    and, second, we appear to need to know the energies of all of states.

    How ever, we have alread y seen this is no t necessarily the case. Since the

    populations (n ,)depend on e-&iikT, ny terms in the

    expansion of

    the

    sum

    for which ii>> kT have near-zero values of

    ni

    can be neglected. This

    30

  • 8/17/2019 Keith a McLauchlan-Molecular Physical Chemistry_ a Concise Introduction-Royal Society of Chemistry (2004)

    43/140

    Partition Functions

    31

    limits the nu mb er tha t needs be considered and has far-reaching conse-

    quences.

    A common approximation is to assume that the various modes of

    energy that a molecule m ay possess are indepe ndent,

    so

    th at for each level

    ‘total = Eelectronic ‘vibrational + ‘rotational + ‘translational

    (2.2)

    This is the sam e assum ption m ade in interpreting molecular spectra. It is

    a very good approximation but it is one and there are many cases in

    which it can be seen to be (slightly) inaccurate.

    A

    simple exam ple is tha t a

    change in the v ibrational energy of a n an harm onic oscillator (e.g. a real

    diatom ic molecule) involves a ju m p to a h igher vibratio nal level in which

    the ave rage internuclear distance differs from th at in the lowest one, and

    therefore causes a change in the moment of inertia and the rotational

    energy of the m olecule, too. H ow ever, the app roxim ation is sufficiently

    good for us to be able to understand the properties of molecules and to

    calculate their thermodynamic properties, often within measurement

    accuracy. If we require their exact values we have no alternative to

    me asuring the energy levels experimentally an d e ntering them into the

    un-approx imated energy equation, without assuming tha t the individual

    contributions are indepe ndent.

    Summ ing over the know n energy levels is the mo st convenient way of

    obtainin